Você está na página 1de 19

Journal of Structural Geology 33 (2011) 500e518

Contents lists available at ScienceDirect

Journal of Structural Geology


journal homepage: www.elsevier.com/locate/jsg

EBSD microfabric study of pre-Cambrian deformations recorded in quartz pebbles


from the Sierra de la Demanda (N Spain)
B. balos a, *, P. Puelles a, S. Fernndez-Armas b, F. Sarrionandia a
a
b

Departamento de Geodinmica, Universidad del Pas Vasco, PO Box 644, E-48080 Bilbao, Spain
Servicios Generales de Investigacin-SGIKER, Universidad del Pas Vasco, PO Box 644, E-48080 Bilbao, Spain

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 29 July 2010
Received in revised form
28 December 2010
Accepted 8 January 2011
Available online 19 January 2011

We describe a new method for the reorientation of lattice preferred orientation data in the absence of
a pre-constrained kinematic reference frame. The method enables us to present conventional quartz
fabric diagrams after measurements taken from rock sections with a general orientation with respect to
foliation and lineation. A microstructural and Electron Back-Scattered Diffraction (EBSD) study of quartz
pebbles in early Cambrian conglomerates following this method permit us to recognize a variety of
fabrics that resulted from syn-metamorphic ductile deformation under variable temperatures up to
650  C. The likely source area of the conglomerates was a Proterozoic basement. Candidates for source
rock correlations include Neoproterozoic units similar to those outcropping in the northern Iberian
Massif, Neoproterozoic medium- to high-grade metamorphic rocks as those outcropping in SW Iberia, or
a Neoarchean to Mesoproterozoic concealed basement.
2011 Elsevier Ltd. All rights reserved.

Keywords:
Quartz
Petrofabric
EBSD
Proterozoic
Cambrian
Iberian Ranges
Spain

1. Introduction
The relationships between sedimentary Cambrian successions
and older rock units have traditionally been a matter of active
research in the geological literature (Van Hise and Leith, 1909). This
has been also the case for the Iberian Peninsula (Vidal et al., 1994),
where the analysis has often been hampered by the paucity of
relevant Rosetta-stone (Hatcher, 1995) outcrops. Conglomerate
pebbles from unconformable formations often carry a memory of
the precedent tectono-sedimentary history, encapsulated in structures and microfabrics that originally were formed in their source
rocks. Structural analysis of such pebble fabrics is usually precluded
by the difculty, or impossibility, to unravel a clear external structural reference framework to which refer any petrographic observation or petrofabric measurement following the conventional
procedures (Passchier and Trouw, 1996). Rock fabrics are related to
second-rank tensorial physical quantities and thus are intrinsic
attributes independent of the external reference system. A corollary
to this is that the perspective of microstructures and three-dimensional crystallographic orientations provided by any two-dimensional section of a rock is thoroughly constrained by its petrofabric.

* Corresponding author. Tel.: 34 4 6012628; fax: 34 4 6013500.


E-mail address: benito.abalos@ehu.es (B. balos).
0191-8141/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsg.2011.01.005

Accordingly, the former might also be used to unravel the latter and
show it in the conventional fabric representations, provided that
pertinent spatial rotations are accomplished.
Outcrops of unconformable Cambrian sediments in contact with
Proterozoic rocks in the northwestern Iberian Massif show that an
angular unconformity separates them and that the latter were
regionally deformed, locally intruded by igneous suites, and mildly
metamorphosed before the Cambrian (de Sitter, 1961; Julivert and
Martnez Garca, 1967; Matte, 1967; Marcos, 1973; Prez Estan,
1973; Martnez Cataln, 1985; Daz Garca, 2006). In correlatable
outcrops of the central Iberian Ranges (Colchen, 1974; lvaro and
Vennin, 1998; Lin et al., 2002) various authors also unravelled
the unconformable relationships between the Cambrian and the
Neoproterozoic (Schmidt-Thom, 1973; Lin and Tejero, 1988).
balos (2001) and lvaro et al. (2008) reported microstructural and
metamorphic features postdated by the lowermost Cambrian rocks
that permit recognition of the Cadomian orogeny (late Neoproterozoic to early Cambrian) in the area.
It is thought that magmatic arcs and back-arc basins are the
Cadomian tectonic settings preserved in the Iberian Peninsula
(Ugidos et al., 1997; Fernndez Surez et al., 1998). The Proterozoic
basement source for the Neoproterozoic sedimentary basin does
not outcrop and very little is known on its lithological nature (Vidal
et al., 1994). Ugidos et al. (2010) reported a remarkable geochemical
homogeneity in the Upper Neoproterozoic series of central and

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

northwestern Spain, likely suggesting an extensive and homogenized source region. Other Proterozoic constituents of Iberia
include granulite-facies allochthonous blocks (interpreted as lower
crustal fragments) recovered from deep submarine exploration in
the thinned continental margins of the Bay of Biscay and in the
Galicia Bank (Gardien et al., 2000). The earliest Cambrian successions cropping out in the core of the Asturian arc and in the Iberian
Ranges contain conglomerate layers that might preserve relics of
such a concealed basement.
In this article we describe a new method for the reorientation of
lattice preferred orientation data in the absence of a pre-constrained kinematic reference frame. The method enables us to
present conventional quartz fabric diagrams measured in rock

501

sections with a general orientation with respect to foliation and


lineation. We apply the technique to investigate the fabric of ductilely deformed quartz pebbles contained by an early Cambrian
conglomerate of North Spain and discuss its implications in the
Proterozoic geology of the Iberian Peninsula.
2. Geological context
2.1. Regional geology
The Sierra de la Demanda of N Spain forms the northwestern
extension of the Iberian Ranges (Fig. 1). This massif contains thick
Cambro-Ordovician successions correlated with coeval and

Fig. 1. Geological sketch map (after Ramrez Merino et al., 1990) of the area to the South of Anguiano and (inset) regional geological context of the Sierra de la Demanda massif. Inset
shows pre-Mesozoic outcrop distribution in the Iberian Peninsula.

502

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

sedimentologically similar series of the West-Asturian-Leonese


Zone of the Iberian Massif (Carls, 1983; Gozalo and Lin, 1988). The
early Paleozoic rocks were deformed and mildly metamorphosed
during the Variscan orogeny (Colchen, 1974) and were unconformably overlain by unmetamorphosed Triassic siliciclastics of the
germano-type Buntsandstein facies, as well as by younger
successions. Tertiary (Alpine) inversion tectonism is manifested
in the area by N-verging major thrusts and associated folds, as well
as by variably intense ductile strain (Liesa and Casas Sainz, 1994; Gil
Imaz, 2001; Capote et al., 2002).
The occurrence of Cambrian rocks that overlie azoic sediments
in the Sierra de la Demanda (N Spain) was rst realized by Schriel
(1930) and afterwards by Lotze (1961). Colchen (1974) disclosed the
stratigraphic organization of these rocks, which was rened by
Palacios (1982) and Shergold et al. (1983). The largest outcrops of

Early Cambrian rocks occur in the eastern part of the Sierra de la


Demanda, to the South of Anguiano (La Rioja), dissected by the
Najerilla River (Fig. 1). These are pre-trilobitic siliciclastic successions containing a basal (up to 300 m thick) unfoliated quartz-rich
conglomerate formation (Anguiano Conglomerate) overlain by
sandstones and quartzites (Areniscas del Puntn). The lowermost
conglomerate beds are stratigraphically 600 m below the oldest
trilobite-bearing Cambrian beds (Palacios, 1982; Palacios and Vidal,
1992; Lin et al., 1993, 2002).
Neoproterozoic slates (Anguiano Schists; cf. Lotze, 1957;
Colchen, 1974) crop out under the Anguiano Conglomerate
(Fig. 1). These are azoic, dark siltstones and mudstones with
interbedded sandstone layers. This unit has a maximum thickness
of 100e150 m, estimated from surface geology (Ramrez Merino
et al., 1990). A Neoproterozoic age is attributed to this succession

Fig. 2. A: View of the Najerilla River Valley near Anguiano (looking toward the South) showing the geological units cited in the text. B: Conglomerate layers interbedded with
sandstones in the upper part of the Anguiano Conglomerate. C: Thick conglomerate beds from the lower part of the Anguiano Conglomerate. D: Close view of a conglomerate
bed with cm-sized, rounded quartzose pebbles. E, F: Field photograph (E) and sketch (F) showing the unconformity between gently (early Cambrian Anguiano Conglomerate) and
steeply dipping (pre-Cambrian Anguiano Schist) formations. Site location is shown in A.

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

based upon regional correlations with similar units cropping out


about 120 km to the SE (Paracuellos Group; Lin and Tejero,
1988; Valladares et al., 2002a). Gently dipping bedding planes are
variably overprinted by a steep foliation associated to subhorizontal
intersection lineations and crenulation microfold axes (balos,
2001).

2.2. The Anguiano Conglomerate


This is an up to 300 m thick siliciclastic succession made of
minor sandstone intercalations and of conglomerates (Figs. 1 and 2)
with more than 90% of vein quartz, chert or quartzite pebbles (a
quartz-pebble conglomerate after Cox et al., 2002). The sedimentary structures and the internal organization in sequences
suggest a conglomeratic beach sedimentary environment. Bed
dips are moderate and no penetrative foliations are observed
(Fig. 2AeD). Lotze (1961) suggested that the contact between
Cambrian and pre-Cambrian successions is unconformable, whereas
Colchen (1974) interpreted that it appears always tectonized. The
basal, unconformable contact of the conglomerate is shown in the
Fig. 2EeF.
This unit contains abundant, well-rounded silicic pebbles mm to
cm in grainsize cemented by diagenetic quartz (e.g.: Figs. 2D and
3AeD). They correspond to ve main groups: (I; Fig. 3A) metamorphic polycrystalline quartz aggregates, (II; Fig. 3B) monocrystalline quartz, (III; Fig. 3E) black cherts, (IV) zircon and
tourmaline-bearing sandstones and (V; Fig. 3F) foliated quartzites.
The undeformed matrix is made of much smaller (diagenetically
overgrown) monocrystalline quartz grains (Fig. 3C) and primary
porosity was diagenetically lled with quartzose precipitates
(Fig. 3D). The matrix includes minor amounts of accessory zircon,
tourmaline, rutile and mica. Pressure-solution microstructures can
be observed locally. They are shown by small pebbles with ellipsoidal contours partly truncated by the straighter traces of polycrystalline quartz aggregates (Fig. 3A) or by matrix quartz grains
that bulge into and dissolve larger monocrystalline quartz pebbles
(Fig. 3B). They are also observable as matrix quartz grain syntaxial
overgrowths preserving ghost inclusions that delineate the
rounded shape of the original detrital grains (Fig. 3C). Likely, these
microstructures formed early during the diagenesis. Cementation
conferred the conglomerates their massive constitution and
mechanically competent nature, preserving them from any subsequent strain or foliation development.
Black chert pebbles (Fig. 3EeF) exhibit microstructural evidence
of brittle deformation (microfaults, brecciations and thin veins lled by syntaxial brous quartz) and are less rounded than other
pebbles. They can exhibit ductile deformations, too (microfolds
with thickened hinges and axial planar continuous foliations;
Fig. 3E). Lin and Tejero (1988) dened a 4 m thick laminated
chert unit with lenticular shales (Fresno Formation) in the Neoproterozoic outcrops of the central Iberian Ranges. However, no
evidence of tectonic deformation of this unit prior to deposition of
Cambrian successions has been reported. Quartzite and sandstone
pebbles often show continuous foliations (Fig. 3F). Sandstone
pebbles are made of quartz and less abundant feldspar grains
embedded by a ner grained matrix with oriented mica and chlorite. These pebbles often contain detrital tourmaline, rutile and
zircon grains (Fig. 3GeH).
Mono- and poly-crystalline quartz pebbles exhibit a variety of
microstructures that resulted from dynamic recrystallization,
a thermally activated process contemporaneous with deformation
in the dislocation creep regime. The concomitant, syn-metamorphic ductile deformations include continuous foliations, quartz
grain shape fabrics, various types of subgrain or recrystallized new

503

grain microstructures, and lattice preferred orientations (LPOs).


They are described in further detail in the following sections.
3. Methods
3.1. Current understanding of quartz LPOs
Quartz microstructures and fabrics resulting from ductile
deformation are well known in ductilely deformed quartz-tectonites. The deformation regime and active intracrystalline slip
systems constrain fabric internal and external symmetry (Bouchez
et al., 1983; Price, 1985; Schmid and Casey, 1986; Passchier and
Trouw, 1996, and references therein). The interpretations derived
from microstructural and fabric studies relate systematically to
oriented thin sections cut along directions normal to the
macroscopic foliation (the XY structural plane) that contain the
lineation (the X structural direction). In the Fig. 4 the basic
geometrical relationships between the XYZ reference framework
and quartz [c] and hai crystallographic axis orientations are shown
schematically for non-coaxial deformations under increasing
temperatures, represented in the conventional XZ section of
stereographic diagrams.
In the Fig. 5 we summarize the geometrical [c] and
hai crystallographic axis organizations that result from non-coaxial
ductile deformation of quartz under temperatures characteristic of
low- to medium-grade metamorphism (Fig. 5AeC) and under highgrade metamorphism (Fig. 5DeF). In both situations quartz fabrics
permit to identify two perpendicular girdles that contain axis
submaxima at specic locations (Fig. 5A and D). As regards the [c]
axes, submaxima can occur along the Y structural direction as well
as near Z or X, whereas three hai axis submaxima separated 120
occur either parallel to the plane normal to the foliation (Fig. 5AeB)
or along a plane normal to both the [c] axis girdle and to the XZ
plane (Fig. 5DeE). The hai axis maxima are often of different
intensity. Depending upon the temperature of deformation, the
most intense submaxima can be located close to X (Fig. 5AeB) or to
Z (Fig. 5DeE). Under non-coaxial deformations, the [c] axis girdles
are oblique to the XYZ reference frame, the angle of this fabric
obliquity being constrained by the intensity of non-coaxial deformation and by the relative contribution of coaxial components
(subsimple shear; cf. Means et al., 1980).
3.2. Analytical methods
Preparation of oriented thin sections directly from quartz
pebbles is not possible in the Anguiano Conglomerate since they
are mechanically coupled to the matrix and no viable chemical or
mechanical separation method is devised. Pebble sections used for
the fabric study have, thus, a priori unknown relationships with
respect to the conventional XYZ structural framework. Only the
foliation trace can be recognized.
Quartz microstructures and fabrics were studied in standard,
30 mm thick polished rock sections. Preliminary studies were performed with a polarizing microscope and a ve-axis Leitz universal
stage (U-stage). An Electron Back-Scattered Diffraction (EBSD)
study was performed on selected thin sections. In order to remove
the mechanically damaged uppermost layer, these were ultra-polished using a colloidal silica suspension for 3e5 h. Samples were
covered with an Au-Pa layer before polishing completion and
nally carbon coated to prevent charging.
Crystallographic orientation measurements were performed at
the University of the Basque Country (Faculty of Science and
Technology) with an automated electron back-scattered diffraction
system (Channel5, HKL) attached to a JEOL JSM-7000F Field Emission Scanning Electronic Microscope (FE-SEM). Samples were

504

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

505

subgrains and new grains (Trimby et al., 1998). The crystallographic


orientation data were thus ltered so that the actual fabric
diagrams of any selected pebble contain one orientation for each
grain. To this end, a critical misorientation threshold of 10 was
considered (allowing boundary completion down to 5 ). Fabric
diagrams are presented in lower hemisphere, equal area stereographic projections where the projection planes lack any conventional geometric relationship with respect to the structural XYZ
framework of the pebbles. The Orientation Distribution Function
(ODF) was calculated from individual orientations with specic
Channel5 software packages using a Gaussian bell curve smoothing
width of 20 . The strength of the texture was expressed by the
texture index J, calculated as the mean square value of the ODF.
3.3. Fabric rotation procedures
In order to present our fabric study in conventional stereographic diagrams (lower hemisphere projections where the foliation appears vertical oriented E-W and the lineation is horizontal
E-W), we envisaged two rotation procedures: one to be applied to
U-stage measurements and the second for the EBSD data. In the
rst case the procedure is based upon the statistical geometric
relationships between the orientation of grain boundaries and the
XZ structural plane. In the second case the procedure takes into
account the geometrical relationships that exhibit quartz [c] (also
[0001]) and hai axes (these including the [11-20] and [10-10]
directions) with respect to foliation and lineation (Figs. 4 and 5).
Quartz [c] axis can be notated as [0001] or, alternatively (this is the
standard notation used by the Channel5 and HKL software), as the
pole to the family of planes {0001}, whereas the hai axes can also be
notated as poles to the sets of m-planes {11-20} or {10-10}.

Fig. 4. Idealized lower hemisphere stereographic projections showing the relationships between quartz [c] and hai crystallographic axis fabrics and intracrystalline slip
systems operating under increasing temperatures and non-coaxial deformation (after
Schmid and Casey, 1986).

mounted in this apparatus on a stage tilted 70 , with the thin


section long axis parallel to the FE-SEM X-axis. The beam working
distance was 20 mm (Prior et al., 1999). An acceleration voltage of
25 kV, a beam current of ca. 14 nA and a copper tape attached to the
sample surface surrounding the measurement area were used to
reduce charging effects. Crystallographic orientation solutions with
Mean Angular Deviation (MAD) values over 1.2 (between detected
and simulated patterns) were rejected to assure EBSD measurement reliability.
Orientation maps and lattice preferred orientation (LPO) pole
gures were drawn using the Channel5 software package after
automated EBSD analysis on predened sampling grid steps of
10 mm. These steps are sometimes signicantly smaller than the
average grainsize of the quartz aggregates and can create artifacts
in the pole gure contours. In order to avoid this, we used the
software package to isolate the areas of the thin sections occupied
by quartz pebbles and, then, to perform automated recognition of

3.3.1. Rotation of U-stage data


As determination of quartz hai axis orientations with a U-stage is
not possible, nding the rotation axes and angles needed to fabric
restoration to the conventional microtectonic XZ framework is
precluded unless various prerequisites are met. We accomplished
U-stage fabric rotations with two pebbles whose quartz microstructure resembled closely well-known examples from the literature. This implied that, whilst a priori the orientation of the XYZ
reference frame was general with respect to the edges and plane of
the thin sections, likely the X and Z axes were close to be contained
in the plane of the pebble thin section, in which the foliation was
clearly observable, too. The rotations performed were sequential
and included rst a vertical axis rotation in order to place the
foliation EW in the stereoplots. Then rotations on horizontal E-W
and N-S axes enabled us to relocate the foliation vertical E-W with
X horizontal. Determination of the vertical axis rotation angle is
straightforward from visual inspection of the angle between the
pebble foliation and the U-stage E-W referential. The tilt angle
needed to reorientate the foliation (XY) perpendicular to the
horizontal U-stage plane was determined from the statistical mean
of the angles formed between the plane of the thin section and
several tens of quartz subgrain boundaries (parallel to prism
planes) oriented parallel to the foliation. In a conventional XZ
section of a rock specimen this set of grain boundaries is expected

Fig. 3. Micrographs showing textural features of the Anguiano Conglomerate. A: Polycrystalline quartz aggregate pebble showing an internal foliation absent in the surrounding
matrix and a pressure-solution contact (marked by the arrow) against a monocrystalline quartz pebble whose original, curved outline appears truncated. B: Small matrix quartz
grains bulging into and dissolving (arrow) a larger monocrystalline quartz pebble. C: Diagenetic quartz overgrowths of two monocrystalline quartz grains whose original rounded
outlines are still depicted by ghost inclusions (marked by the arrows). D: Quartzose diagenetic precipitates with rhythmic and idiomorphic arrangement lling the primary
porosity. E: Black chert pebble showing microfolds, an axial planar continuous foliation and an extensional vein normal to it. F: Ellipsoidal quartzite pebble (left) with an internal
foliation adjacent to an irregular black chert pebble (right) containing thin quartz-lled cracks and microbreccia. G: Detrital matrix tourmaline grain containing a zircon inclusion
that has developed a radioactive damage halo. H: Detrital matrix zircon grain with a slightly rounded contour and internal idiomorphic crystal faces denoting magmatic growth
textures. Plane polarized light in D and H and crossed polars in the other micrographs.

506

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

Fig. 5. A: Conventional quartz [c] and hai axis LPOs (equal area, lower hemisphere) resulting from non-coaxial ductile deformation under low- to medium-grade metamorphic
temperatures. B: Geometrical relationships between girdle and point maxima intersections with respect to the structural XYZ referential. C: The foliation strike, the principal quartz
hai axis point maxima and the [c]-girdle/hai-girdle intersection can be determined directly from LPOs, independent of the internal or external reference framework considered. D:
Conventional quartz [c] and hai axis LPOs resulting from non-coaxial ductile deformation under high-grade metamorphic temperatures. E: Idem as B. F: Idem as C. See text for
further details.

to be statistically perpendicular to the plane of the section as well


as parallel to the foliation. Similarly, the tilt angle needed to reorientate the X structural direction to a horizontal E-W position was
determined with the U-stage from the statistical mean of the angles
formed between the plane of the thin section and several tens of
subgrain boundaries oriented normal to the foliation.
3.3.2. Rotation of EBSD data
In this case the foliation strike (observable in pebble sections)
and the hai and [c] girdles recognizble in EBSD fabric diagrams
permitted to constrain the orientation of the XY plane. As depicted
by the Fig. 5BeC and EeF, the foliation and the [c] crystallographic
axis girdles always intersect parallel to the Y structural direction.
Such intersection can be marked directly (Fig. 5DeE; the intersection coincides with Y) or indirectly (Fig. 5AeB; the intersection
occurs at 90 to Y) by the hai and [c] girdles. Once this is achieved,
the determination of the Z and X directions is straightforward (Z is
normal to the XY plane and X is normal to Y and Z within XY; Fig. 5B
and E).
The next step corresponds to rotation of both the crystallographic fabrics measured and their XYZ reference axes to the
conventional microtectonic stereographic framework, that consists
of a vertical XY plane with X horizontal E-W. Since the EBSD fabric
analysis provides complete crystallographic orientation data for
quartz hai and [c] axis girdles and the Channel5 software package
requires Euler angles to handle rotations, we actually followed with
the EBSD data the procedure described in the Fig. 6. The quartz
EBSD fabrics obtained for each pebble sliced with a general orientation permitted to determine hai and [c] axes best-t great circles
after the stereographic projections. These intersect at a point easy
to nd (Fig. 6A), as well as is the strike of the foliation (straightforward from inspection of the microstructure).

In low- to moderate-temperature quartz deformation (Fig. 5AeC)


the hai and [c] axes girdle intersection is at 90 from the Y structural
axis, which is also contained by the [c] axis girdle (Fig. 6B). The
foliation strike and the orientation of Y are then used to establish the
orientation of the XY plane and its pole (Fig. 6C). The pole to XY is the
Z structural direction (Fig. 6D), which forms an angle (4, the fabric
obliquity angle) with the hai and [c] girdle intersection when plastic
deformation includes rotational components.
In a similar way, in high-temperature quartz deformation
(Fig. 5DeF) the hai and [c] axes girdles intersect along Y. The foliation strike and the orientation of Y can be used to dene the
orientation of the XY plane, its pole (the Z structural direction) and
the X direction (at 90 of Y within XY). Z forms the 4 fabric obliquity
angle with the pole to the [c] girdle and 904 with the pole to the
hai girdle.
Determination of the three Euler angles (a, b and g) needed to
perform rotations requires consideration of orientation of the
current X, Y and Z structural axes together with the nal position of
these reference axes used in conventional fabric stereoplots (Bunge,
1985, pp. 79e80): horizontal E-W (X), N-S (Z) and vertical (Y). The
stereoplots drawn by the Channel5 software after raw data use as
reference axes three directions: the horizontal E-W (H(EW)EBSD in
the Fig. 6E), the horizontal N-S (H(NS)EBSD in the Fig. 6E) and the
vertical (VEBSD in the Fig. 6E) which relate to the edges of the
rectangular rock thin sections used for measurements. These EBSD
reference axes and the crystallographic axes must be sequentially
rotated using the Euler angles so that the actual stereoplots can be
interpreted conventionally. Determination of the Euler angles a,
b and g requires to locate rst the line of nodes (Fig. 6E), which is
parallel to the intersection between the XZ structural plane and the
plane containing the horizontal axes of the EBSD referential. On this
basis, a is the angle between the line of nodes and the X structural

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

507

Fig. 6. Lower hemisphere stereographic projections explaining the procedure followed to rotate EBSD measured raw quartz LPOs. A: The quartz [c] and hai LPOs measured in
pebbles whose structural reference framework is unknown are used to identify major circles that t best the girdle or point maxima distributions. B: The Y structural axis is located
along the [c] axis girdle, 90 apart from the [c] and hai girdle intersection. C: The location of Y and the strike of the foliation are used to reconstruct the orientation of the pebble
foliation (XY structural plane) and its pole. D: The Z structural direction coincides with the pole to the foliation and the X direction is normal to both Y and Z. The mist between Z
and the [c] and hai girdle intersection is the fabric obliquity angle. E: The referential for EBSD measurements and the XYZ structural reference are related through the Euler angles
(see text for further details). F: Once the Euler angles are determined, the original quartz [c] and hai crystallographic axis fabrics can be rotated so that X appears oriented horizontal
East-West, Z horizontal normal to X, and Y vertical, as in conventional fabric diagrams.

axis, whereas b is the angle between the Y structural axis and the
vertical EBSD referential and, nally, g is the angle between the line
of nodes and the horizontal E-W EBSD referential. The sign
convention followed was the described by Ramsay and Huber
(1983, p. 168). Various sets of Euler angles, each related to one of
the quartz pebble fabrics measured, were used as inputs in the
Channel5 software to recalculate and redraw the fabric stereoplots
as conventional fabric diagrams (Fig. 6F). The results are presented
and interpreted in the following section 4.
4. Quartz pebble syn-metamorphic fabrics
4.1. Microstructures and metamorphic grade
Microstructures and fabrics indicating syn-metamorphic ductile
deformations can be recognized in several pebble sections. Foliations in polycrystalline quartz pebbles are dened by the parallel
orientation of grain boundaries in combination with shape fabrics
parallel or oblique to them (Fig. 3A, G, and H). In polymineralic
aggregates (sandstones, quartzites and cherts) foliations are
dened by the parallel orientation of graphite, mica and chlorite as
well as by the elongation of quartz grains. These foliations are in
some lithologies axial planar with microfolds showing slight limb
thinning and hinge thickenning (Fig. 3H).
The mineral assemblages preserved in these pebbles denote
very low-grade or low-temperature greenschists facies metamorphism. Polycrystalline quartz aggregates lack mineral assemblages appropriate to identify any metamorphic grade. However,
they exhibit a considerable variety of microstructures resulting

from ductile deformation (Fig. 7AeC, 8A-C and 9A-G). Microstructures pointing to dislocation creep (undulose extinction,
deformation lamellae, subgrains with boundaries parallel to prism
planes) and grain boundary migration (recrystallized new grains,
shape fabrics, oriented mineral inclusions) are common. They
characterize Hirth and Tullis (1992) regime 1 of quartz plastic
ow, which operates under temperatures below 250e300  C and/
or fast strain rates. Also present are microstructures denoting
dislocation climb and recovery (subgrain formation, attened
grains), which characterize Hirth and Tullis (1992) regime 2 of
quartz plastic ow. These reect deformation under higher
temperature, slower strain rates and probably in the presence of
a catalyzing uid phase. Finally, mosaic microstructures and shape
fabrics denote a complete dynamic recrystallization (through
either subgrain rotation or grain boundary migration). The high
mobility of grain boundaries is demonstrated here by the
complete inclusion of oriented chlorite crystals by recrystallized
quartz grains. These microstructures characterize Hirth and Tullis
(1992) regime 3 of quartz plastic ow, which occurs under higher
temperatures (often above the transition between the greenschists
and the amphibolite metamorphic facies) and slower strain rates
than the regime 2.
4.2. High-temperature quartz fabrics
Some of the microstructures observed correspond to the
chessboard microstructures of quartz (Bouchez et al., 1985),
regarded as a microstructural geothermometer (Kruhl, 1996) that
denotes plastic deformation under medium to high-grade

508

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

Fig. 7. A: Micrograph of a polycrystalline quartz pebble (crossed polars) with chessboard microstructure and a continuous foliation. B: Detail of A showing the statistical
parallelism (or perpendicularity) between quartz grain boundaries and the foliation. C: Micrograph of the eld of view B made with a compensation gypsum lens to highlight the
preferred crystallographic orientation of the microstructure. D: Sketch of the conventional thin section that contains the pebble shown in A. E: Lower hemisphere, equal area
stereographic projection of 200 quartz [c] axes measured from the thin section D and plotted taking its long and short edges as E-W and N-S referentials. F. Lower hemisphere, equal
area stereographic projection of the 200 quartz [c] axes rotated so as to present the foliation vertical E-W. See text for further details.

metamorphic conditions and possibly under hydrous conditions


and fast strain rates (Mainprice et al., 1985). They occur in pebbles
with a prominent foliation and several quartz grains. When
observed with crossed polars under the petrographic microscope,
grains remain close to extinction while the stage is rotated 360 .
This means that their [c] axes are normal to the plane of pebble
sections, which thus are close to the orientation of the XZ structural
plane. Chessboard subgrain and grain boundaries are parallel to
prismatic and basal crystallographic planes (Tuba and Cuevas,
1985) due to prism-hai and prism-[c] intracrystalline slip in the
high-quartz stability eld (Kruhl, 1996). This highlights that ductile
deformation was constrained by intracrystalline slip system transitions active at moderate to elevated temperature (above
550e600  C under natural strain rates; Okudaira et al., 1995).
In the Figs. 7 and 8 we present the results of microfabric analysis
and restoration of two polycrystalline quartz pebbles with the
microstructure described above. Quartz grain shape fabrics are
evident in both pebbles (Figs. 7A and 8A) and the chessboard
arrangement of quartz grain boundaries is obvious in the Fig. 8A
(and to a lesser extent in parts of the Fig. 7A). Petrographic observations with the gypsum plate permit to demonstrate that the
quartz shape fabrics are related to crystallographic preferred
orientations (Figs. 7BeC and 8BeC). The respective quartz [c] axis

fabrics were determined with the U-stage, originally taking the thin
section long and short edges (Figs. 7D and 8D) as horizontal E-W
and N-S referentials. They are presented as raw stereoplots (with
a sample reference frame) in the Figs. 7E and 8E, where the great
circles represent foliation orientations determined after its strike in
the thin sections and the mean tilt angle of grain boundaries
parallel to that direction. Fabric rotations following the procedure
described in section 3.3.1 resulted in the stereoplots presented in
the Figs. 7F and 8F. In the Fig. 7F two perpendicular [c] axis girdles
(the type II fabrics of Lister and Dornsiepen, 1982) intersect at
a point close to the stereonet center and are oblique to the foliation
and its pole. This pattern suggests that the structural X direction is
close to a horizontal E-W line and that, thus, the stereonet horizontal plane actually coincides with the XZ structural section. It
denotes activation of prism [c] slip in addition to the basal hai and
[c] intracrystalline slip systems. In the quartz pebble of Fig. 8
preservation of several grain boundaries perpendicular to the foliation enabled to determine the orientation of the X structural
direction in addition to that of the foliation and its pole. This
permitted to accomplish a complete rotation of the original fabric
(Fig. 8E) to an XZ stereoplot (Fig. 8F). The most prominent feature of
the restored fabric is a single [c] axis point maxima close to X and
oblique to the XYZ framework. It denotes the high-temperature

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

509

Fig. 8. See caption of Fig. 7. In F. the pebble foliation is vertical E-W and the quartz mineral lineation is horizontal within the foliation plane. See text for further details.

activation of prism [c] slip (at T of ca. 650  C), as well as the noncoaxial character of ductile deformation.
The results of an EBSD-based study of subgrain and new grain
misorientations associated to chessboard microstructures are presented in the Fig. 9. Micrographs 9A and 9C show two sample
microstructures with the superposed tracks (running from the
upper left to the lower right) of two straight lines along which
quartz lattice orientations were recorded at steps of 10 mm. In the
line graphs to the right (Fig. 9B and D) the misorientation angle of
the quartz lattice at each track pixel (measured in  ) with respect to
the orientation of the adjoining one (the abscise refers to track
distance in mm) is plotted. The misorientation angles correspond to
rotations along specic, rationale crystallographic axes that can be
labeled with the Channel5 software. The Fig. 9B line graph shows
quartz grain boundaries that are systematically associated to 60
reorientations around the [0001] axis of the bounded lattices.
Actually, these are quartz basal plane-parallel grain boundaries
related to prism-[c] slip, hardly detectable with the petrographic
microscope (Tuba and Cuevas, 1985). The 60 lattice rotation effect
around the [0001] axis is likely due to Dauphin twinning. Brazil
twin planes, which are normal to quartz hai axes, also separate right
and left hand quartz lattices misoriented 60 , but are not detected
by the EBSD technique. In order for Dauphin twins to be produced

by dislocations, they must be twist boundaries parallel with the


basal plane. This requires dislocations slipping in the basal plane
with more than one Burgers vector. Tullis and Tullis (1972) showed
that Dauphin twinning contributes to the total quartz fabric, as it
may inuence the relative activities of competing slip systems
(Lister, 1979; Lister and Paterson, 1979). It operates independently
of intracrystalline slip and does not reorient the [c] axis. The Fig. 9D
line graph depicts misorientations below 1 (only a few times
above 2 ) along a large and non-uniform set of rationale directions
for most parts of the track. In general, these can be considered
orientation mists below the measuring technique resolution and
thus are disregardable.
4.3. Moderate- and low-temperature quartz deformation
The fabric of various pebbles made of polycrystalline quartz
aggregates was studied with the EBSD technique (Fig. 10AeG).
They exhibit a foliation and a dynamically recrystallized microstructure. The original crystallographic orientation data were
measured at points regularly separated 10 mm and ltered
before plotting them in the stereoplots. Channel5 software was
used to discriminate between grains with misorientations above
a critical threshold of 10 , allowing boundary completion down

510

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

Fig. 9. A: EBSD orientation map of a chessboard quartz microstructure. Quartz crystallographic orientation along the transect 1e2 was recorded at steps of 10 mm. B: Plot of the
quartz lattice misorientation angle between adjacent measurement points along the track 1e2 shown in A. C: Idem as A except that the misorientation prole is shown in D. D: Idem
as B. See text for further details.

to 5 . A single crystallographic orientation was assigned to each


grain and plotted in the stereograms. The fabrics obtained often
show point and girdle axis distributions with general orientations with respect to the horizontal and vertical reference axes
of the EBSD acquisition. The values of the texture index J vary
between 1.18 and 4.46, thus characterizing moderate to intense
fabrics.
In the Fig. 11 the quartz [c] and hai LPOs are plotted in
a horizontal XZ reference plane after the rotation procedure
described in section 3.3.2 The monoclinic distribution of the [c]
axes with respect to the XYZ framework is evident, denoting the
non-coaxial character of quartz plastic deformation. The [c] axes
appear arranged in point or girdle distributions. In the former
case, they always contain point maxima concentrations along the
Y axis, which are either isolated or combined with other
concentrations close to Z or at positions between Y and Z. Except
for the case of Fig. 11C, axis distributions correspond to incomplete girdles partly constrained by the three point maxima
orientations already reported. Fabric asymmetry is not so
straightforward from inspection of the hai axes distributions.
These are spread out along girdles close to the horizontal
projection plane of the stereoplots that can contain point submaxima concentrations evenly spaced 60 . The features described
can be interpreted as a result of the operation of the prismhai intracrystalline slip system in all the fabrics, in addition to
basal hai slip in some cases (Fig. 11AeC) and rhomb-hai slip in
others (Fig. 11D, E and G). This unravels quartz plastic deformation under moderate temperatures (Fig. 4) of the greenschists or
low-T amphibolite metamorphic facies. The quartz fabric of
Fig. 11E presents an incomplete [c] axis girdle forming a low angle
with the foliation, thus resembling operation of prism-[c] slip
under higher deformation temperatures. The distribution of the
hai axes along the periphery of the steroplot, however, precludes

such an interpretation, since the pattern expected for prism-[c]


slip is a girdle normal to the X direction (Fig. 4).
The monocrystalline quartz pebble group includes pebbles of
variable size in which petrographic inspection enables to recognize
that they are made either of a single quartz crystal or of a small
number of crystals with a close orientation (subgrains). The second
were studied with the EBSD technique (Fig. 12). The original crystallographic orientation data set was ltered with the Channel5
software before plotting in the stereographs by considering
misorientations above 5e10 (as described above). The fabrics
obtained contain a single [c] axis point maxima normal to hai axis
girdles made of three point maxima separated 60 among them.
The strength of these fabrics is remarkable according to the texture
index J, which attains values of up to 12.0. Low and high-angle grain
boundaries are widespread and exhibit complex, sutured geometries. Distinction between old grains and recrystallized subgrains is
not evident always. The variation of [c] or hai axis orientations
within each pebble attains 20 with respect to the orientation of
the maximum density concentrations. This gure clearly exceeds
the limit used to discriminate dynamically recrystallized new
grains reported above, but should be interpreted taking into
account that it does not relate to subgrains in direct contact
misoriented up to 20 (which would imply they are new grains
instead of subgrains) but to sets of subgrains (from the subgrain
population of a given pebble) that were progressively and accumulatively misoriented relative to each other. These microfabric
and textural features enable us to interpret that the monocrystalline quartz pebbles endured cold working under temperatures lower than those inferred for the plastic deformation of the
polycrystalline pebbles.
Some quartz pebbles that resemble polycrystalline aggregates
because they exhibit a foliation and a recrystallized microstructure
produced crystallographic fabrics similar to those of the

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

511

512

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

monocrystalline quartz pebbles (Fig. 13). The texture index J of their


LPOs, however, varies between 1.34 and 3.82. The fabrics contain
single [c] axis point maxima normal to hai axis girdles made of
three ill dened point maxima separated 60 . The variation of the
[c] or hai axis orientations within each pebble usually exceeds the
20 reported in the precedent paragraph with respect to the
orientation of the maximum density concentrations, and even
reaches 30e40 . A more intense, protracted or hotter plastic
deformation of former quartz monocrystals can be inferred for
these pebbles, but interpretations on the active intracrystalline slip
systems cannot be made with condence after the data available.
4.4. Microchert pebbles and conglomerate matrix fabrics
Black chert pebbles exhibit petrographic evidence of either
brittle or ductile deformation. The small size of quartz grains makes
it impossible to ascertain from petrographic inspection if they
underwent crystal plasticity or any other deformation mechanism
(volume diffusion, solution-precipitation). Our EBSD study shows
that microcherts lack any crystallographic preferred orientation
(Fig. 14A and B), the texture index J being very low. This suggests
that ductile deformation was assisted by mechanisms different
from the crystal plasticity reported for quartz pebbles. However, it
does not exclude a comparable tectonic setting (under low to
moderate temperature conditions) prevalent during their folding
and foliation development. Trepmann et al. (2010) have shown that
dissolution-precipitation creep at low differential stress can be
a deformation mechanism operative under low-grade metamorphic conditions (of up to 350  C) promoting foliation development in quartzose rocks.
The matrix quartz aggregates (Fig. 14C) lack any crystallographic
preferred orientation. The preservation of intact diagenetic microtextures and the lack of macroscopic or petrographic evidence of
ductile strain suggests that the conglomerates behaved as
mechanically brittle, competent units during most of their Paleozoic burial history after deposition and during subsequent Variscan
and Alpine deformations.
5. Discussion
The Anguiano Conglomerate is an outstanding example of
quartz-pebble conglomerate. This rock type is thought to represent tectonically quiescent conditions under which weathering
(chemical and mechanical) was very efcient and included protracted transport, prolonged mechanical abrasion, intense
chemical weathering or recycling of older conglomerate (Cox
et al., 2002, and references therein). Its occurrence in the
geological record increases backward through time, most examples being Proterozoic. Cox et al. (2002) suggested that this rock
forms by postdepositional diagenetic alteration (clast disintegration/dissolution along pebble boundaries, porosity collapse by
compaction and cementation) of polymict conglomerate precursors. Therefore, in spite of being made of more than 90% of quartz,
chert or quartzite clasts, it is usually derived from lithological
diverse sources (Grange et al., 2010). Petrographic evidence
described in the section 2.2 of this study demonstrates quartz
dissolution from pebbles and precipitation in the conglomerate
matrix (as syntaxial overgrowths of sand-sized quartz grains and
primary porosity llings). The only evidence of a polymict

conglomerate composition consists of accessory minerals such as


tourmaline, zircon, rutile and scarce mica, as well as of the
various quartzose pebble lithologies (metachert, sandstone,
quartzite and vein quartz). Any other original pebble composition,
had it really existed, would have been disintegrated by postdepositional diagenetic alteration.
The textural characteristics of the Anguiano Conglomerate
pebbles and the relationships with their matrix/cement permit to
interpret that they preserve microfabrics originally formed in the
source rocks and that, thus, pebble quartz plastic deformation was
acquired during stages of their tectonic history preceding erosion
and early Cambrian deposition. Our petrographic, U-stage and EBSD
fabric study, together with the fabric restoration method used,
enabled us to identify deformations produced under temperatures
characteristic of low-, medium- and high-grade metamorphism. The
original tectonic setting was an orogenic environment that actually
exposed in the pre-early Cambrian Earths surface rocks that had
been buried previously to variable depths. It also contained nonmetamorphic silicic metasediments and a broad variety of recrystallized quartz veins, likely of igneous afnity. The occurrence of
detrital zircon grains with internal magmatic microstructures in
metasandstone pebbles and in the matrix point to the occurrence of
magmatic precursors in the source areas. Detrital tourmaline is often
common in those pebbles and the matrix, but this mineral might be
related either to magmatic precursors or to metamorphic/hydrothermal recycling of preexisting tourmaline-bearing rocks, boronrich sediments or even marine evaporites (Dini et al., 2008; PrezXavier et al., 2008; Torres-Ruiz et al., 2003). Bea et al. (2009) have
highlighted the potential of this mineral to date orogenic deformation phases. Rutile associated to zircon relics in foliated metasandstone pebbles point toward medium- to high-grade metamorphic or
plutonic source rocks.
Andersen and Picard (1971) discussed how quartz pebble types
had previously been used in provenance studies. This question, as
well as those discussed above, bear on the characterization of
currently concealed sediment provenance regions. Garzanti et al.
(2007) distinguished ve types of orogenic domains as the
primary building blocks of composite orogens, each producing
predictable detrital modes, heavy mineral assemblages and
unroong trends. Their axial belt (the metamorphic backbone of
collisional orogens) and, to a lesser extent, magmatic arc source
regions are the best candidates in our case study. According to
Ugidos et al. (2010, and references therein), in the case of the
Iberian realm these regions would correspond to reworked
orogenic belts and internal cratons of peri-Gondwanan afnity
whose remnants are preserved in Neoproterozoic to early Paleozoic
sedimentary successions.
The lithological and accessory mineral evidence preserved in
the Anguiano Conglomerate suggests that magmatic, low- to highgrade metamorphic and sedimentary rocks conformed the
conglomerate source area. It might be reasonably expected that
the outcrops of those source rocks would not occur too far from the
current conglomerate outcrops. Source candidates include three
tectonic settings: (1) the outcropping Neoproterozoic domains of
the Iberian Massif (located toward the West), (2) the poorly known
mid- and lower crustal segments of the Iberian microplate continental margin (Gardien et al., 2000), and (3) the concealed Ebro
Massif (located toward the East), currently hindered under a kmthick Mesozoic to Tertiary cover.

Fig. 10. EBSD maps (labeled A-G, located to the left) and raw crystallographic fabrics of polycrystalline quartz pebbles. The surrounding matrix is displayed with a lighter color
palette. Equal area, lower hemisphere stereographic projections with density color patterns and added contour lines in multiples of uniform distribution. Min and Max
correspond to the minimum and maximum density concentration values. J is the texture index. Crystallographic notation: the poles of the family of planes {0001} are the quartz
[c] axes, whereas the poles to the sets of planes {11-20} or {10-10} correspond to the hai axes. See text for further details. (For interpretation of the references to color in this gure
legend, the reader is referred to the web version of this article).

Fig. 11. Quartz LPOs for the polycrystalline quartz pebbles shown in the Fig. 10, plotted in a horizontal XZ reference plane after the rotation procedure described in the Fig. 6 and the
manuscript. See caption to Fig. 10 and the text for further details.

Fig. 12. EBSD maps (labeled AeF and located to the left of the gure) and raw crystallographic fabrics of various monocrystalline quartz pebbles. See caption to the Fig. 10 for further
details.

Fig. 13. EBSD maps (labeled AeF and located to the left of the gure) and raw crystallographic fabrics of various quartz pebbles apparently resembling polycrystalline aggregates.
See caption to the Fig. 10 for further details.

516

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

Fig. 14. EBSD maps (labeled AeC and located to the left of the gure) and raw crystallographic fabrics for black chert pebbles (A and B) and the quartzose matrix among pebbles (C).
See caption to the Fig. 10 for further details.

As regards the rst case, low- to high-grade, NeoProterozoic


metamorphic rocks showing high-temperature ductile quartz
deformation are only known to occur in the Ossa-Morena Zone
(balos, 1992; balos and Eguluz, 1992), a peri-Gondwanan
magmatic arc accreted to the Iberian autochthon during the
Cadomian orogeny (balos et al., 2002 and references therein).
These rocks exhibit a remarkable and widespread Neoproterozoic
magmatic arc afnity that is not evident in the Sierra de la
Demanda. lvaro and Vennin (1998) reported the occurrence of
schist pebbles (in addition to quartzite, metasediment and black
quartzite ones) in early Cambrian conglomerates of the central
Iberian Ranges. These constitute the Daroca Formation, which is
slightly younger than the Anguiano Conglomerate. The authors
interpreted that the source area of the pebbles contained igneous
and metamorphic rocks also compatible with an axial belt or
magmatic arc tectonic setting. As regards the second source
candidate, so far it is known to exist in the thinned continental
margins of the Bay of Biscay and the Galicia Bank (Gardien et al.,
2000), though its autochthonous setting is unknown. Radiometric
datings permitted to identify them as Neoarchean to Paleoproterozoic metamorphic rocks, some of which were already exhumed in
the Mesoproterozoic. Finally, the geological constitution of the Ebro
Massif remains largely unknown, as to our knowledge it has never
been drilled and its existence is suggested by stratigraphic studies
on the source regions of terrigenous Mesozoic successions
outcropping in the Iberian Ranges and parts of the Basque Cantabrian Basin.
The wealth of radiometric ages recovered from inherited detrital
zircon grains from different parts of the Iberian Massif have

provided accurate time constraints on the source regions of the


Neoproterozoic sedimentary rocks (Fernndez-Surez et al., 1999,
2000, 2002, 2003; Gutirrez Alonso et al., 2005; Abati et al.,
2010). These relate notably to Neoproterozoic (Pan-African and
Cadomian) orogens, and to a lesser extent to Paleoproterozoic
(1.8e2.1 Ga) or Neoarchean (2.4e2.8 Ga) ones, that resemble the
West-African craton (e.g.: Ugidos et al., 1997; Valladares et al.,
2002b). Mesoproterozoic relics were found in units of exotic
origin (allochthonous) with respect to the Iberian Massif and in
autochthon, Neoproterozoic to early Paleozoic sedimentary
successions (e.g.: Fernndez Surez et al., 2003; Gutirrez Alonso
et al., 2005).
We argue that the source area of the Anguiano Conglomerate, with pebbles that preserve medium- to high-grade
metamorphic fabrics, would correspond to current mid- to
lower crustal settings of the Iberian microplate. These might
either be related to the concealed Ebro Massif or to crustal
segments buried under the Neoproterozoic successions of the
Iberian Massif. The radiometric datings of high-grade metamorphic rocks recovered from the continental margins of the
Bay of Biscay suggest that some of them (of Neoarchean to
Paleoproterozoic age) were already exhumed in the Mesoproterozoic (Gardien et al., 2000). Neoproterozoic orogenic
processes are better documented in the Iberian Peninsula and
might have provided source metamorphic tectonite components for the Anguiano Conglomerate, too. Combined radiometric and metamorphic grade studies of recycled tectonite
relics would be still more efcient to constrain provenance
studies.

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

6. Conclusions
Conglomerate pebbles encapsulate structures and microfabrics
that carry a memory of tectonic activity older than the age of the
hosting sedimentary formations. Structural analysis of such pebble
fabrics is precluded by the difculty, or impossibility, to unravel any
external structural reference framework to which refer petrographic observations or petrofabric measurements following the
conventional procedures. The pebbles are sawed with a general,
statistically unspecied orientation with respect to their internal
foliation and lineation and, under this circumstance, microstructural observations are of limited value. Except in special cases, most
crystallographic preferred orientation measurements would be
difcult to interpret in terms of intracrystalline slip systems and
deformation temperature or regime. These shortcomings can be
resolved with the fabric rotation methods described in this study,
applied to quartz LPO data acquired with the U-stage and the EBSD.
The quartz microstructures and the crystallographic fabrics
presented in this article permit to constrain the kinematic framework of pre-early Cambrian tectonite pebbles and unravel the slip
systems active during their ductile deformation. This gave us a way
to identify deformations produced under temperatures characteristic of low-, medium- and high-grade metamorphism. Source rock
candidates for the Anguiano conglomerate include Neoproterozoic
units as those outcropping in the northern Iberian Massif, Neoproterozoic medium- to high-grade metamorphic rocks as those of
SW Iberia, or a poorly known Neoarchean to Mesoproterozoic
basement concealed under the Iberian Phanerozoic successions.
Acknowledgments
The scientic comments of O. Fernandez, M. Pearce, G. Zulauf
and an anonymous reviwer, and the editorial handling of J. Hippert,
helped to improve the quality of the original manuscript and are
sincerely acknowledged. Financial support was provided by the
Spanish Ministerio de Ciencia e Innovacin (Grupo Consolidado,
project CGL2008-01130/BTE) and the Universidad del Pas Vasco
(project GIU09/61).
References
balos, B., 1992. Variscan shear-zone deformation of late Precambrian basement in
SW Iberia, implications for circum-Atlantic pre-Mesozoic tectonics. Journal of
Structural Geology 14, 807e823. doi:10.1016/0191-8141(92)90042-U.
balos, B., 2001. Nuevos datos microestructurales sobre la existencia de deformaciones precmbricas en la Sierra de la Demanda (Cordillera Ibrica). Geogaceta
30, 3e6.
balos, B., Eguluz, L., 1992. Structural geology of the Mina Afortunada Gneiss Dome
(Badajoz-Crdoba shear zone, SW Spain). Annales Tectonic 6, 95e110.
balos, B., Carreras, J., Druguet, E., Escuder Viruete, S., Gmez Pugnaire, M.T., Lorenzo lvarez, S., Quesada, C., Rodrguez Fernndez, L.R., Gil Ibarguchi, J.I., 2002.
Variscan and pre-Variscan tectonics. In: Gibbons, W., Moreno, T. (Eds.), The
Geology of Spain. Geological Society, London, pp. 156e183.
Abati, J., Gerdes, A., Fernndez Surez, J., Arenas, R., Whitehouse, M.J., Dez
Fernndez, R., 2010. Magmatism and early-Variscan continental subduction in
the northern Gondwana margin recorded in zircons from the basal units of
Galicia, NW Spain. Geological Society of America Bulletin 122, 219e235.
doi:10.1130/B26572.1.
lvaro, J.J., Vennin, E., 1998. Stratigraphic signature of a terminal Early Cambrian
regressive event in the Iberian Peninsula. Canadian Journal of Earth Sciences 35,
402e411.
lvaro, J.J., Bauluz, B., Gil-Imaz, A., Simn, J.L., 2008. Multidisciplinary constraints on
the Cadomian compression and early Cambrian extension in the Iberian Chains,
NE Spain. Tectonophysics 461, 215e227. doi:10.1016/j.tecto.2008.04.006.
Andersen, D.W., Picard, M.D., 1971. Quartz extinction in siltstone. Geological Society
of America Bulletin 82, 181e186.
Bea, F., Pesquera, A., Montero, P., Torres-Ruiz, J., Gil Crespo, P.P., 2009. Tourmaline
40
Ar/39Ar chronology of tourmaline-rich rocks from Central Iberia dates the
main Variscan deformation phases. Geologica Acta 7, 399e412. doi:10.1344/
104.000001446.
Bouchez, J.L., Lister, G.S., Nicolas, A., 1983. Fabric asymmetry and shear sense in
movement zones. Geologische Rundschau 72, 401e419.

517

Bouchez, J.L., Tuba, J.M., Mainprice, D., 1985. Dformation naturelle du quartz:
coexistence des systmes de glissement de direction -a- et -c- haute
temperature (migmatites de la nappe dOjn, Espagne). Comptes Rendus de
lAcadmie des Sciences de Paris 301, 841e846.
Bunge, H.J., 1985. Representation of preferred orientations. In: Wenk, H.-R. (Ed.),
Preferred Orientation in Deformed Metals and Rocks: An Introduction to
Modern Texture Analysis. Ed. Academic Press, Orlando, pp. 73e108.
Capote, R., Muoz, J.A., Simn, J.L., Liesa, C.L., Arlegui, L.E., 2002. Alpine tectonics I:
the Alpine system north of the Betic Cordillera. In: Gibbons, W., Moreno, T.
(Eds.), The Geology of Spain. Geological Society, London, pp. 367e400.
Carls, P., 1983. La Zona Asturoccidental-Leonesa en Aragn y el macizo del Ebro
como prolongacin del Macizo Cantbrico. In: Comba, JA (Ed.), Contribuciones
sobre temas generales. Libro Jubilar J.M. Ros, v. 3. Instituto Geolgico y Minero
de Espaa, Madrid, pp. 11e32.
Colchen, M., 1974. Gologie de la Sierra de la Demanda (Burgos-Logroo, Espagne).
Memorias del Instituto Geolgico y Minero de Espaa 85, 1e436.
Cox, R., Gutman, E.D., Hines, P., 2002. Diagenetic origin for quartz-pebble
conglomerates. Geology 30, 323e326.
de Sitter, L.U., 1961. Le Prcambrien dans la Chane cantabrique. Comptes Rendus
Sommaire des Sances de la Societ Gologique de France 9, 1e253.
Daz Garca, F., 2006. Geometry and regional signicance of Neoproterozoic
(Cadomian) structures of the Narcea Antiform, NW Spain. Journal of the
Geological Society, London 163, 499e508.
Dini, A., Mazzarini, F., Musumeci, G., Rocchi, S., 2008. Multiple hydro-fracturing by
boron-rich uids in the Late Miocene contact aureole of eastern Elba Island
(Tuscany, Italy). Terra Nova 20, 318e326. doi:10.1111/j.1365-3121.2008.00823.x.
Fernndez Surez, J., Daz Garca, F., Jeffries, T.E., Arenas, R., Abati, J., 2003.
Constraints on the provenance of the uppermost allochthonous terrane of the
NW Iberian Massif: inferences from detrital zircon U-Pb ages. Terra Nova 15,
138e144. doi:10.1046/j.1365-3121.2003.00479.X.
Fernndez Surez, J., Gutirrez Alonso, G., Jeffries, T.E., 2002. The importance of
along-margin terrane transport in northern Gondwana: insights from detrital
zircon parentage in Neoproterozoic rocks from Iberia and Brittany. Earth and
Planetary Science Letters 204, 75e88.
Fernndez Surez, J., Gutirrez Alonso, G., Jenner, G.A., Jackson, S., 1998. Geochronology and geochemistry of the Pola de Allande granitoids (northern Spain).
Their bearing on the Cadomian/Avalonian evolution of NW Iberia. Canadian
Journal of Earth Sciences 35, 1e15.
Fernndez Surez, J., Gutirrez Alonso, G., Jenner, G.A., Tubrett, M.N., 1999. Crustal
sources in lower Paleozoic rocks from NW Iberia: insights from laser ablation UPb ages of detrital zircons. Journal of the Geological Society, London 156,
1065e1068.
Fernndez Surez, J., Gutirrez Alonso, G., Jenner, G.A., Tubrett, M.N., 2000. New
ideas on the Proterozoic-Early Paleozoic evolution of NW Iberia: insights from
U-Pb detrital zircon ages. Precambrian Research 102, 185e206.
Gardien, V., Arnaud, N., Desmurs, L., 2000. Petrology and Ar-Ar dating of granulites
from the Galicia Bank (Spain): African craton relics in Western Europe. Geodinamica Acta 13, 103e117.
Garzanti, E., Doglioni, C., Vezzoli, G., And, S., 2007. Orogenic belts and orogenic
sediment provenance. Journal of Geology 115, 315e334.
Gil Imaz, A., 2001. La estructura de la Sierra de Cameros: deformacin dctil y su
signicado a escala cortical. Instituto de Estudios Riojanos, Logroo. Ciencias de
la Tierra, 23, 1e305 (Ph.D. Thesis, Univ. Zaragoza, 1999).
Gozalo, R., Lin, E., 1988. Los materiales hercnicos de la Cordillera Ibrica en el
contexto del Macizo Ibrico. Estudios Geolgicos 44, 399e404.
Grange, M.L., Wilde, S.A., Nemchin, A.A., Pidgeon, R.T., 2010. Proterozoic events
recorded in quartzite cobbles at Jack Hills, Western Australia: new constraints
on sedimentation and source of > 4 Ga zircons. Earth and Planetary Science
Letters 292, 158e169. doi:10.1016/j.epsl.2010.01.031.
Gutirrez Alonso, G., Fernndez Surez, J., Collins, A.S., Abad, I., Nieto, F., 2005.
Amazonian Mesoproterozoic basement in the core of the Ibero-Armorican Arc:
40Ar/39Ar detrital mica ages complement the zircons tale. Geology 33,
637e640. doi:10.1130/G21485.1.
Hatcher Jr., R.D., 1995. Structural Geology, Principles, Concepts and Problems,
second ed. Prentice-Hall, Upper Saddle River, New Jersey, p. 525.
Hirth, G., Tullis, J., 1992. Dislocation creep regimes in quartz aggregates. Journal of
Structural Geology 14, 145e159.
Julivert, M., Martnez Garca, E., 1967. Sobre el contacto entre el Precmbrico y el
Cmbrico en la parte meridional de la cordillera Cantbrica y el papel del
Precmbrico en la orognesis Herciniana. Acta Geolgica Hispnica 2, 107e110.
Kruhl, J.H., 1996. Prism- and basal-plane parallel subgrain boundaries in quartz:
a microstructural geothermobarometer. Journal of Metamorphic Geology 14,
581e589.
Liesa, C.L., Casas Sainz, A.M., 1994. Reactivacin alpina de pliegues y fallas del
zcalo hercnico de la Cordillera Ibrica: ejemplos de la Sierra de la Demanda y
la Serrana de Cuenca. Cuadernos del Laboratorio Geolgico de Laxe 19,
119e135.
Lin, E., Tejero, R., 1988. Las formaciones precmbricas del antiforme de Paracuellos (Cadenas Ibricas). Boletn de la Real Sociedad Espaola de Historia
Natural 84, 39e49.
Lin, E., Perejn, A., Sdzuy, K., 1993. The loweremiddle Cambrian stages and stratotypes from the Iberian Peninsula: a revision. Geological Magazine 130, 817e833.
Lin, E., Gozalo, R., Palacios, T., Gmed-Vintaned, J.A., Ugidos, J.M., Mayoral, E.,
2002. Cambrian. In: Gibbons, W., Moreno, T. (Eds.), The Geology of Spain.
Geological Society, London, pp. 17e29.

518

B. balos et al. / Journal of Structural Geology 33 (2011) 500e518

Lister, G.S., 1979. Fabric transitions in plastically deformed quartzites: competition


between basal, prism and rhomb systems. Bulletin de Minralogie 102, 232e241.
Lister, G.S., Dornsiepen, U.F., 1982. Fabric transitions in the Saxony granulite terrain.
Journal of Structural Geology 4, 81e92.
Lister, G.S., Paterson, M.S., 1979. The simulation of fabric development during plastic
deformation and its application to quartzite: fabric transitions. Journal of
Structural Geology 1, 99e115.
Lotze, F., 1957. Zum Alter nordwestpanischer Quartzit-Sandstein-Folgen. Neues
Jahrbuch fr Geologie und Palontologie Monatshefte 10, 128e139.
Lotze, F., 1961. Das Kambrium Spaniens. Teil I: Stratigraphie. Abhandlungen der
Gesellschaft der Wissenschaften zu Gttingen (Mathematisch-Physikalische
Klasse) 6, 283e501.
Mainprice, D., Bouchez, J.L., Blumeeld, P., Tuba, J.M., 1985. Dominat c-slip in
naturally deformed quartz: implications for dramatic plastic softening at high
temperature. Geology 14, 819e822.
Marcos, A., 1973. Las series del Paleozoico inferior y la estructura Herciniana del
occidente de Asturias (NO de Espaa). Trabajos de Geologa 6, 1e113.
Martnez Cataln, J.R. 1985. Estratigrafa y estructura del Domo de Lugo (sector
Oeste de la Zona Asturoccidental-Leonesa). Corpus Geologicum Gallaeciae 2,
1e291 (Ph.D. Thesis Univ. Salamanca, 1981).
Matte, Ph, 1967. Le Prcambrien suprieur schisto-grseux de lOuest des Asturies
(Nord-Est de lEspagne) et ses relations avec les sries prcambriennes plus
internes de larc galicien. Comptes Rendus de lAcadmie des Sciences de Paris
264, 1769e1772.
Means, W.D., Hobbs, B.E., Lister, G., Williams, P.F., 1980. Vorticity and non-coaxiality
in progressive deformations. Journal of Structural Geology 2, 371e378.
Okudaira, T., Takeshita, T., Hara, I., Ando, J., 1995. A new estimate of the conditions
for transition from basal hai to prism [c] slip in naturally deformed quartz.
Tectonophysics 250, 31e46.
Palacios, T., 1982. El Cmbrico entre Viniegra de Abajo y Mansilla (Sierra de la
Demanda, Logroo). Trilobites e ichnofsiles. Publicaciones del Instituto de
Estudios Riojanos, pp. 1e86.
Palacios, T., Vidal, G., 1992. Lower Cambrian acritarchs from northern Spain: the
Precambrian-Cambrian boundary and biostratigraphic implications. Geological
Magazine 129, 421e436.
Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics. Springer, Berlin-Heidelberg.
p. 289.
Prez Estan, A., 1973. Datos sobre la sucesin estratigrca del Precmbrico y la
estructura del extremo Sur del Antiforme del Narcea (NW de Espaa). Breviora
Geologica Asturica 17, 5e16.
Prez-Xavier, R., Wiedenbeck, M., Trumbull, R.B., Dreher, A.M., Monteiro, L.V.S.,
Rhede, D., de Araujo, C.E.G., Torresi, I., 2008. Tourmaline B-isotopes ngerprint
marine evaporites as the source of high-salinity ore uids in iron oxide coppergold deposits, Carajs Mineral Province (Brazil). Geology 36, 743e746.
doi:10.1130/G24841A.1.
Price, G.P., 1985. Preferred orientations in quartzite. In: Wenk, H.-R. (Ed.), Preferred
Orientation in Deformed Metals and Rocks: An Introduction to Modern Texture
Analysis. Ed. Academic Press, Orlando, pp. 385e406.
Prior, D.J., Boyle, A.P., Brenker, F., Cheadle, M.C., Day, A., Lopez, G., Potts, G.J.,
Reddy, S., Spiess, R., Timms, N., Trimby, P., Wheeler, J., Zetterstrom, L., 1999. The
application of electron backscatter diffraction and orientation contrast imaging
in the SEM to textural problems in rocks. American Mineralogist 84, 1741e1759.

Ramrez Merino, J.I., Oliv Dav, A., lvaro Lpez, M., Hernndez Samaniego, A.,
1990. Mapa y Memoria de la Hoja n 241 (Anguiano) del Mapa Geolgico de
Espaa a escala 1:50.000 (Serie MAGNA). Instituto Geolgico y Minero de
Espaa, Madrid.
Ramsay, J.G., Huber, M.I., 1983. Strain Analysis. In: The Techniques of Modern
Structural Geology, vol 1. Academic Press, London, pp. 1e307.
Schmid, S.M., Casey, M., 1986. Complete fabric analysis of some commonly observed
quartz C-axis patterns. In: Mineral and Rock Deformation: Laboratory Studies.
The Paterson Volume, vol 36. American Geophysical Union, Geophysical
Monograph, pp. 263e286.
Schmidt-Thom, M., 1973. Beitrge zur Feinstratigraphie des Unterkambriums in
den Iberischen Ketten (Nordst-Spanien). Geologisches Jahrbuch, Reihe B7,
3e43.
Schriel, W., 1930. Die Sierra de la Demanda und die Montes Obarenes. Abhandlungen der Gesellschaft der Wissenschaften zu Gttingen (MathematischPhysikalische Klasse). Neue Folge 16, 463e567.
Shergold, J.H., Lin, E., Palacios, T., 1983. Late Cambrian trilobites from the Najerilla
formation, north-eastern Spain. Palaeontology 26, 71e92.
Torres-Ruiz, J., Pesquera, A., Gil-Crespo, P.P., Velilla, N., 2003. Origin and petrogenetic implications of tourmaline-rich rocks in the Sierra Nevada (Betic Cordillera, southeastern Spain). Chemical Geology 197, 55e86.
Trepmann, C.A., Lenze, A., Stckhert, B., 2010. Static recrystallization of vein quartz
pebbles in a high-pressure e low-temperature metamorphic conglomerate.
Journal of Structural Geology 32, 202e215. doi:10.1016/j.jsg.2009.11.005.
Trimby, P.W., Prior, D.J., Wheeler, J., 1998. Grain boundary hierarchy development in
a quartz mylonite. Journal of Structural Geology 20, 917e935.
Tuba, J.M., Cuevas, J., 1985. Fbrica del cuarzo en tectonitas de alta temperatura
(Manto de Ojn, Cordilleras Bticas). Estudios Geolgicos 41, 147e155.
Tullis, J., Tullis, T.E., 1972. Preferred orientation produced by mechanical Dauphin
twinning: thermodynamics and axial experiments. American Geophysical
Union. Monograph 16, 67e82.
Ugidos, J.M., Snchez-Santos, J.M., Barba, P., Valladares, M.I., 2010. Upper Neoproterozoic series in the Central Iberian, Cantabrian and West Asturian Leonese
Zones (Spain): geochemical data and statistical results as evidence for a shared
homogenised source area. Precambrian Research 178, 51e58. doi:10.1016/
j.precamres.2010.01.009.
Ugidos, J.M., Armenteros, I., Barba, P., Valladares, M.I., Colmenero, J.R., 1997.
Geochemistry and petrology of recycled orogen-derived sediments: a case
study from Upper Precambrian siliciclastic rocks of the Central Iberian Zone,
Iberian Massif, Spain. Precambrian Research 84, 163e180.
Valladares, M.I., Barba, P., Ugidos, J.M., 2002a. Precambrian. In: Gibbons, W.,
Moreno, T. (Eds.), The Geology of Spain. Geological Society, London, pp. 7e16.
Valladares, M.I., Ugidos, J.M., Barba, P., Colmenero, J.R., 2002b. Contrasting
geochemical features of the Central Iberian Zone shales (Iberian Massif, Spain):
implications for the evolution of Neoproterozoic-lower Cambrian sediments
and their sources in other peri-Gondwanan areas. Tectonophysics 352, 121e132.
Van Hise, C.R., Leith, C.K., 1909. Pre-Cambrian Geology of North America. United
States Geological Survey Bulletin n 360. Government Printing Ofce, Washington, pp. 1e939.
Vidal, G., Palacios, T., Gmez Vintaned, J.A., Dez Balda, M.A., Grant, S.W.F., 1994.
Neoproterozoic-early Cambrian geology and palaeontology of Iberia. Geological
Magazine 131, 729e765.

Você também pode gostar